Articles

CIRCUMBINARY CHAOS: USING PLUTO'S NEWEST MOON TO CONSTRAIN THE MASSES OF NIX AND HYDRA

, , and

Published 2012 July 23 © 2012. The American Astronomical Society. All rights reserved.
, , Citation Andrew N. Youdin et al 2012 ApJ 755 17 DOI 10.1088/0004-637X/755/1/17

0004-637X/755/1/17

ABSTRACT

The Pluto system provides a unique local laboratory for the study of binaries with multiple low-mass companions. In this paper, we study the orbital stability of P4, the most recently discovered moon in the Pluto system. This newfound companion orbits near the plane of the Pluto–Charon (PC) binary, roughly halfway between the two minor moons Nix and Hydra. We use a suite of few body integrations to constrain the masses of Nix and Hydra, and the orbital parameters of P4. For the system to remain stable over the age of the solar system, the masses of Nix and Hydra likely do not exceed 5 × 1016 kg and 9 × 1016 kg, respectively. These upper limits assume a fixed mass ratio between Nix and Hydra at the value implied by their median optical brightness. Our study finds that stability is more sensitive to their total mass and that a downward revision of Charon's eccentricity (from our adopted value of 0.0035) is unlikely to significantly affect our conclusions. Our upper limits are an order of magnitude below existing astrometric limits on the masses of Nix and Hydra. For a density at least that of ice, the albedos of Nix and Hydra would exceed 0.3. This constraint implies they are icy, as predicted by giant impact models. Even with these low masses, P4 only remains stable if its eccentricity e ≲ 0.02. The 5:1 commensurability with Charon is particularly unstable, combining stability constraints with the observed mean motion places the preferred orbit for P4 just exterior to the 5:1 resonance. These predictions will be tested when the New Horizons satellite visits Pluto. Based on the results for the PC system, we expect that circumbinary, multi-planet systems will be more widely spaced than their singleton counterparts. Further, circumbinary exoplanets close to the three-body stability boundary, such as those found by Kepler, are less likely to have other companions nearby.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

New Horizons has one more object to study when it visits Pluto in 2015 (Stern 2008; Young & Stern 2010). Hubble Space Telescope (HST) imaging recently revealed a faint moon orbiting in the plane of the Pluto–Charon (PC) binary (Showalter et al. 2011, S11). The moon, temporarily named P4, orbits between the previously known outer moons Nix and Hydra, which are 10 times brighter (Weaver et al. 2006).

Even before the discovery of P4, the Pluto system was dynamically intriguing. The orbital periods of Charon, Nix, and Hydra are very close to a 1:4:6 ratio (Buie et al. 2006). Despite thorough searching, no resonant lock has been identified (Tholen et al. 2008, T08). The close proximity to resonance begs for an explanation (as discussed further in the conclusions). The stakes are raised further by the proximity of P4 to a 5:1 commensurability with Charon (S11).

Efforts to understand these complex interactions are hindered by the difficulty of measuring the masses of Nix and Hydra. The tightest constraints come from the T08 orbit solution, which fits astrometric data with four body integrations. However, the low masses of Nix and Hydra make their dynamical perturbations difficult to measure. The T08 fits restrict Nix and Hydra to mass ratios below ∼10−4 with Pluto. These limits allow a wide range of plausible albedos and densities.

The sizes of Nix and Hydra are indirectly constrained by their modest photometric variability, which suggests a roughly spherical shape (Stern et al. 2007, S07). Nix and Hydra would have to be large, with diameters ≳ 130 km, for gravity to naturally make them spherical. Such large sizes are marginally inconsistent with the T08 mass constraints and quite inconsistent with our results. New Horizons should clarify lingering uncertainties about the size and shape of Nix and Hydra (Young et al. 2008).

We exploit the precarious position of P4's orbit as an alternate approach to constraining system masses. The stability of P4 depends both on its uncertain orbital parameters, and the strength of perturbations it receives from Nix and Hydra. We use a series of numerical integrations to constrain simultaneously the masses of Nix and Hydra, and the orbit of P4.

Prior to the announcement of P4, Pires Dos Santos et al. (2011) investigated the stability of test particle orbits in the PC system. Using the T08 masses, they found a pocket of low eccentricity orbits between Nix and Hydra that were stable for 105 Charon periods (1800 yr). Our paper assesses longer term stability (up to 3 × 107 yr) for orbits in the vicinity of P4 for a range of Nix and Hydra masses. Ideally, we would demonstrate stability up to the age of the solar system. However, the short period of Charon (6.38 days) and the need to investigate many trial orbits make longer integrations costly.

Because studying stability in the PC system places constraints on satellite masses, our dynamical studies also provide information on the composition of the bodies. By themselves, masses give reasonable constraints on surface albedos and sizes. Once albedos and sizes are determined by New Horizons, internal densities can be determined. The compositional diversity of bodies in the same system is in turn a powerful input to formation theories (Benecchi et al. 2009; Stern 2009).

As the prototypical—and also the second most massive and best studied—Kuiper Belt object, Pluto represents a critical stage in planet formation. As a transitional object, it hold clues both to the processes that form the first planetesimals in gas disks (Chiang & Youdin 2010; Youdin 2010) and the collisional coagulation and destruction that in some places produces planets and in others produces debris disks (Kenyon & Luu 1999; Kenyon 2002; Kenyon & Bromley 2010; Lambrechts & Johansen 2012).

We begin in Section 2 by describing parameters of the PC system, establishing which parameters are reasonably well constrained, and which must be varied in our simulations. Section 3 describes basic stability considerations for the PC system. Appendix A finds the most circular orbits about the PC binary, a non-trivial task due to the non-Keplerian nature of orbits about a binary (Lee & Peale 2006). Readers interested in the main results can skip to Section 4, which presents our numerical integrations of the Pluto system. In Section 5, we summarize our main findings and discuss their implications for exoplanet studies in Section 5.1. The Appendix addresses technical aspects of the integrations including the initialization of particles on the most circular orbit about a binary (Appendix A) and details of the numerical code (Appendix B).

2. PLUTO SYSTEM PARAMETERS

2.1. Size and Mass of Nix, Hydra, and P4

Nix, Hydra, and P4 are detected in reflected visible light. Their diameters depend on their unknown albedos, A, as

Equation (1a)

Equation (1b)

Equation (1c)

To relate apparent magnitudes to size, we adopt a Charon-like phase coefficient (Buie et al. 1997; T08).

Mass estimates from photometry rely on an assumed density. With ρ1 = ρ/(1 g cm−3), the mass ratios relative to Pluto are

Equation (2a)

Equation (2b)

Equation (2c)

for Nix, Hydra, and P4. To express (without loss of generality) masses in terms of albedos, and not the more cumbersome Aρ−2/31, we assume ρ1 = 1, unless stated otherwise. Neglecting the possibility of high porosity, we also refer to A = ρ1 = 1 as the "minimum mass" case.

The orbit solution of T08 gives μNix = (4.4 ± 3.9) × 10−5 and μHyd = (2.5 ± 4.9) × 10−5. The large uncertainties reflect the difficulty of measuring small mass ratios astrometrically. Since the uncertainties extend toward or beyond zero,1 the T08 solution places upper limits ∼10−4 on the mass ratios of Nix and Hydra. The equivalent albedo constraint is A ≳ 0.04. Our stability calculations place much tighter constraints of A ≳ 0.3 on Nix and Hydra.

2.2. Orbit of P4

The orbit of P4 is only loosely constrained. The discovery (S11) announces P4 as "consistent with" a circular, coplanar orbit. Its mean motion from HST images spanning ∼20 days implies an orbital period of 32.1 ± 0.3 days, corresponding to a period ratio of 5.03 ± 0.05 with Charon.

For a Keplerian orbit about the PC barycenter (a good approximation at the 1% level), the measured period corresponds to a mean separation of 57, 400 ± 400 km. Figure 1 plots this orbit location as a dark green dot with error bars. The lighter green error bars show the larger range of orbits considered in our numerical simulations.

Figure 1.

Figure 1. Dynamical environment of P4 is plotted in the space of semimajor axis, a, and eccentricity, e, relative to the Pluto–Charon center of mass. The green dot (with error bars) gives the approximate location of P4. Vertical lines give nominal positions of mean motion resonances. Red and blue dashed lines give the nominal locations of the first-order resonances with Hydra and Nix, respectively. The black dot-dashed line denotes the 5:1 resonance with Charon (which nearly overlaps the 5:4 resonance with Nix). Black diagonal lines show the Nix and Hydra crossing trajectories, N-X and H-X, respectively. Magenta curves labeled CT plot the critical Tisserand parameter relative to Nix and Hydra. The upper and lower CT curves are for low and high masses of Nix and Hydra, i.e., high and low albedos as labeled. See the text for details.

Standard image High-resolution image

The mean motion is consistent with the measured projected radial separation of 59, 000 ± 2000 km from Pluto (which is 2040 km from the barycenter). The pre-discovery HST images that date back to 2006 (S11), as well as future observations from HST, ALMA, and New Horizons will help constrain the orbit of P4.

2.3. Orbital Ephemerides

As summarized in Table 1, we adopt the orbit solution of T08 for the orbits of Pluto, Charon, Nix, and Hydra and the masses of Pluto and Charon. Ephemerides from JPL's HORIZONS system2 are similar, but not identical to the T08 solution. Orbit integrations started with the JPL (specifically the PLU021) ephemerides gave statistically similar results to the T08 ephemerides. Thus, we only present results based the T08 orbit solution.

Table 1. Keplerian Orbital Elementsa and Cartesian State Vectorsb from Tholen et al. (2008)

  Charon Nix Hydra   Charon Nix Hydra
Mass (kg)c 1.52e21 (5.8e17) (3.2e17) μc 0.1166 (4.4e−5) (2.5e−5)
a (km) 19570.3 49240 65210 x −0.1677 −0.1744 0.5202
e 0.0035 0.0119 0.0078 y −0.2551 0.7148 −0.8822
i 96.168 96.190 96.362 z 0.0 5.6384e−5 3.4397e−3
ω 157.9 244.3 45.4 $\dot{x}$ 1.5995 −1.0282 1.0625
M 237.0 129.80 129.12 $\dot{y}$ −1.0452 −0.3580 0.4458
Ω 223.054 223.202 223.077 $\dot{z}$ 0.0 −3.1683e−3 −1.8596e−4
P (days) 6.3872 25.49 38.85 $P \over P_{\rm PC}$ 1 3.99 6.06

Notes. aPlutocentric for Charon and Barycentric for Nix and Hydra. The Nix and Hydra masses are the T08 values, not those used in this work. bCartesian state vectors are given in code units where G = M = Ppc = 1. The Pluto–Charon orbit has been rotated into the xy plane. cM = 1.304e22 kg, μ is mass relative to Pluto.

Download table as:  ASCIITypeset image

Though the T08 (and also JPL) solutions assume specific masses for Nix and Hydra, we vary the masses of Nix and Hydra without varying the orbit solution. The error introduced is hopefully modest compared to the astrometric uncertainty. Future work could establish how orbit solutions vary with uncertain masses.

Salient features of the system's orbital parameters, including circularity, coplanarity, and the proximity of moons to resonances, are discussed in the introduction and evident in Table 1. Though low, the finite eccentricity of the PC orbit, eC = 3.5 × 10−3, has long been considered suspicious. Tidal interaction between Pluto and Charon damps eC on short, ∼10 Myr, timescales (Dobrovolskis et al. 1997; Lithwick & Wu 2008a). Nix and Hydra cannot excite eC above ∼10−5 (LP06); P4 cannot contribute significantly either. External forcing—from solar and planetary tides, Kuiper Belt Objects (KBO) flybys and collisions—appears unable to explain the observed eccentricity (Stern et al. 2003).

After the submission of this manuscript, a lower eccentricity for Charon was announced (Buie et al. 2012). This new solution gives a 1σ limit of "3 km out of round" (Buie et al. 2012), implying eC ≲ 1.5 × 10−4 (at 1σ). A preliminary investigation shows that setting eC = 0 in the T08 orbit solution has a minor effect on orbital stability (Section 4.4).

3. BASIC STABILITY CONSIDERATIONS

This section summarizes previous results on orbital stability that are relevant to the Pluto system in particular and to multi-satellite systems about a binary in general. While these general considerations cannot predict the stability of P4, they are useful in understanding the dynamical environment of P4 as shown in Figure 1, and in interpreting the numerical simulations in Section 4. Our most significant general conclusion concerns the comparison of previous work on multi-planet stability about a single star (Section 3.3) to our simulation of the PC circumbinary system. The comparison shows that binary perturbations can significantly enhance the destabilizing effects of satellite interactions, which is a small step to a more complete understanding of these complex dynamical interactions.

3.1. Stability of Circumbinary Orbits

The stability of a test particle orbiting an inner binary is well studied (Szebehely 1980; Dvorak 1986), including the systematic numerical investigation of Holman & Wiegert (1999, hereafter HW). The HW results show that Nix, Hydra, and P4 are all individually stable about the PC binary. Based on the results of HW (see their Equation (3) and/or their Table 7), period ratios below 2.8 are unstable in the PC system. Nix's period ratio of ≈4 is 41% larger than the stability boundary, and also avoids a possible instability strip near the 3:1 resonance.

3.2. Stability of Two Satellites

We now temporarily ignore the important fact that Pluto and Charon are a binary and consider the three-body stability of two satellites about a central mass that places the combined mass of Pluto and Charon at their barycenter. This approximate problem allows us to simply demonstrate how strong interactions with Nix or Hydra restrict the allowed orbits of P4, as in Figure 1, and also aids the interpretation of numerical results.

The well-known stability criterion for initially circular orbits with semimajor axes a1 and a2 = a1 + Δ requires Δ > 3.5RH in the restricted three-body problem when one of the satellites is massless (Gladman 1993; which also analyzes two massive satellites). That is, two satellites on circular orbits are stable if their orbits are separated by more than 3.5 Hill radii,

Equation (3)

of the massive satellite.

The separation of P4 from Nix or Hydra in terms of Hill radii depends on the assumed masses for Nix and Hydra. For the mass scalings in Equation (2), P4 is separated from Nix by $29 \sqrt{A_{\rm Nix}}$ and from Hydra by $17 \sqrt{A_{\rm Hyd}}$ Hill radii (of Nix and Hydra, respectively). Separation by fewer than 3.5RH would require either ANix < 0.014 or AHyd < 0.042. Such low albedos and thus high masses are inconsistent with the T08 (1σ) upper mass limits. The more stringent mass limits that we find ensure that P4 is stable by the Hill stability criterion for circular orbits.

The stability of an eccentric test mass follows from the (approximately) conserved Tisserand parameter

Equation (4)

where a, e,  and i (inclination) refer to the test body (P4) and a' refers to the perturber (Nix or Hydra). The stability boundary of circular, coplanar (e = i = 0) orbits at |aa'| = 3.5RH defines a critical CT. (Restricted three-body) stability for arbitrary e and i holds for CT above the critical value.

Figure 1 plots the critical CT curves, sometimes called "Tisserand tails." The outer and inner tails of Nix and Hydra, respectively, are relevant for interactions with P4. For low Nix and Hydra masses (the A = 1 curves), the tails would only cross an eccentric P4, eP4 ≳ 0.1. With higher Nix and Hydra masses (the A = 0.05 curves), Hydra's Tisserand tail intersects plausible P4 orbits at lower eC, especially if P4 lies on the outer edge of allowed orbits. While a considerable simplification, these considerations from the restricted three-body problem demonstrate how higher masses for Nix and Hydra severely restrict the allowed orbits of P4.

3.3. Stability of Three Satellites

Still ignoring perturbations from the central binary, we now consider three interacting satellites about a central mass. Sharp stability boundaries no longer exist, but the timescale to orbit crossing can be studied numerically.

Most investigations of multi-planet stability consider roughly equal mass companions with evenly spaced orbits, in terms of Hill radii. Chambers et al. (1996, hereafter C96) defined the mutual Hill radius, as

Equation (5)

for neighboring planets (i and i + 1). This definition does not reduce to the standard Hill radius when μi → 0. This deficiency is readily corrected by a mass-weighted averaging of the semimajor axes. This distinction is insignificant when the satellite masses are similar, but is relevant here due to the low mass of P4.

C96 measured the orbit crossing timescale, tc, for systems of three equal mass planets, with planet–star mass ratios ranging from μ = 10−9 to μ = 10−5. Relative to the period of the inner planet, P1, we fit the data in Figure 4 of C96 as

Equation (6)

where Δ' is the orbit separation in mutual Hill radii and the functional form is motivated by the Faber & Quillen (2007) study of systems of >3 planets. The mass scaling Δ'μ1/12∝μ−1/4 differs from the μ−1/3 Hill radius scaling due to three-body resonances (Quillen 2011).

Unequal masses (of P4 in particular) make direct application of these results difficult. To allow simple estimates, we make a range of possible assumptions: μ as the average of all three satellites or of just Nix and Hydra; the mutual Hill radius defined as in Equation (5) or with density weighted ai. In all cases, the minimum mass A = 1 case is stable, with crossing times >1030 yr. For the higher mass A = 0.05 case, tc ≲ 3 × 106 yr, implying rapid instability.

Even neglecting binary perturbations, the stability of Pluto's minor moons depends strongly on their assumed masses. Binary perturbations accelerate the destabilization. We show in Section 4 that for the high-mass case with A = 0.05, the simulated crossing time for P4 is ≲ 103 yr, over a thousand times faster than the above estimate neglecting the central binary.

3.4. External Perturbations and Tides

In this paper, we approximate the Pluto system as a set of isolated point masses. Since Pluto's Hill sphere extends ∼120 times further than Hydra's orbit, solar tides are a minor effect. The protective 3:2 resonance between Neptune and Pluto helps to weaken the dominant planetary perturbation. Nevertheless, follow-up work should test the effect of weak external perturbations on long-term stability.

Collisions with interloping KBOs could significantly perturb the weakly bound outer moons. If the Kuiper Belt was massive in its youth, collisions would unbind ∼100 km class binary KBOs (Nesvorný et al. 2011; Parker & Kavelaars 2012). Pluto's moons, P4 in particular, could be destabilized by collisions that only modestly perturb its eccentricity. Combining the dynamical stability of Pluto's moons with collisional perturbations could be a powerful constraint on both the Pluto system and the collisional environment of the Kuiper Belt.

Tides in the Pluto system are dominated by interactions between Pluto and Charon, which are locked in a dual synchronous state, with both spin periods equaling the orbital period (Dobrovolskis et al. 1997). In principle, the eccentricities of minor moons should be damped by exciting the eccentricity of Charon, which then suffers tidal dissipation. However, Lithwick & Wu (2008b) conclude that this damping mechanism is weak.

4. RESULTS FOR LONG-TERM STABILITY

4.1. 4 + N-body Integrations

We study the stability of P4 with a suite of 4 + N-body integrations. In these simulations, the four massive bodies are Pluto, Charon, Nix, and Hydra. Treating P4 as a massless test particle allows simultaneous investigation of many trial orbits. We follow each test particle until it crosses the orbit of either Nix or Hydra. (No significant perturbations to the orbits of the massive bodies occurs in the simulations.) Integrations were performed using the 15th-order Radau integrator (Everhart 1985), as implemented by the Swifter software package.3 Details concerning code performance are deferred to Appendix B.

Between different sets of simulations, we vary the uncertain masses of Nix and Hydra, keeping their mass ratio fixed at MNix/MHyd = 0.575, the value implied by their median optical brightness. This mass ratio assumes that Nix and Hydra share the same density and albedo, which should be expected for standard formation scenarios.4 The range of Nix and Hydra masses considered corresponds (for a density of 1 g cm−3) to albedos from 0.03 to 0.24. Integrations were run with albedos up to 0.4 (μNix  ≈  10−6), however only 10%–30% of test particles suffered orbit crossing after ∼108 Charon orbits, making systematic investigations too expensive.

The initial conditions for the massive bodies are given in Table 1. The initial orbits for the test particles were chosen by two different methods. The first suite of simulations, described in Section 4.2, populated P4 orbits by randomly sampling Keplerian elements. The second suite of simulations, summarized in Section 4.3 initializes P4 on the "most circular" orbits about the PC binary.

4.2. Uniformly Sampled Keplerian Orbits

For each adopted mass of Nix and Hydra, we integrate 5000 P4 orbits with initial conditions chosen randomly from a uniform distribution of Keplerian osculating elements.5 The semimajor axes are restricted to the range

Equation (7)

or a/aPC = 2.89–3.01. Eccentricity and inclination are restricted to e < 0.05 and i < 0fdg5. The modest range in i was insignificant for our results and will not be discussed further. Other Keplerian angles (argument of pericenter, longitude of ascending node, and mean longitude) are sampled over the full (2π) range. Figure 1 compares the sampled orbits to the nominal location of mean motion resonances as well as Nix and Hydra crossing trajectories.

4.2.1. Mapping ae Space

Figure 2 maps the median stability timescale versus initial a and e for several Nix and Hydra masses. As shown by the color bars on each map, crossing times increase significantly as the mass of Nix and Hydra drops from high to low (upper left to lower right plots). Crossing times are always short at higher a and e (upper right of each plot). Interactions with Hydra are the likely culprit, consistent with the Tisserand parameter curves in Figure 1.

Figure 2.

Figure 2. Median crossing time, tc, vs. initial a (as Δa = a − (5 × 104 km)) and e of P4. The mass of Nix and Hydra is labeled as both A (their shared albedo for ρ = 1 g cm−3) and m1 (mass relative to the minimum mass at A = 1). The color scale for the tc is above each panel. Crossing times are longer for lower masses of Nix and Hydra and lower eccentricities of P4. The semimajor axis dependence is more complicated, due to the effect of resonances. See the text for details.

Standard image High-resolution image

At low eccentricity, crossing times are longest, and display a complex dependence on a. Resonances play a key role. Indeed resonance overlap is a generic cause of orbital chaos (Mudryk & Wu 2006). From the approximate resonance locations in Figure 1, specific resonances can be implicated.

The 5:1 resonance with Charon helps destabilize the region near a ∼ 5.8 × 104 km (Δa ∼ 8.0[ × 103 km] as labeled in Figure 2) at lower masses. The 4:5 resonance with Hydra shortens crossing times between 5.65 ≲ a/(104 km) ≲ 5.70 (i.e., 6.5 ≲ Δa/(103 km) ≲ 7.0), for the highest mass (upper left). This change in the prominence of different resonances is expected as the mass of Nix and Hydra varies relative to PC.

4.2.2. Median Crossing Times: Measured and Extrapolated

Figure 3 shows how crossing times scale with the mass of Nix and Hydra. Median timescales are plotted for three subsets of the initial orbital parameters: (1) all initial orbits, (2) only e < 0.015, and (3) the stable "core" of parameters with both e < 0.015 and 5.70 × 105 km < a < 5.75 × 105 km. For the "core" sample, we also plot the 90th percentile of crossing time (beyond which only 10% of particles survive). At any mass, crossing times increase as cuts become more selective. The "missing" data points for low-mass cases occur where P4 orbits were so stable that the median (or 90th percentile) timescale was not reached after 108 PC periods.

Figure 3.

Figure 3. Crossing timescale for P4 vs. the mass of Nix and Hydra. The bottom axis scales masses relative to the albedo one case. The top axis shows the corresponding albedo (for ρ = 1 g cm−3). Circles give the crossing times from simulations. Dashed lines are power-law fits to the mass dependence. Colors denote different sets of initial P4 orbits. From bottom to top, median crossing times are shown for (blue) the full range of orbital parameters, (green) e < 0.015, and (black) a stable "core" with both e < 0.015 and 5.70 × 105 km < a < 5.75 × 105 km. Finally, red data points give the 90th percentile of longest lived orbits in the core. If extrapolation can be trusted, Nix and Hydra require A ≳ 0.5 to ensure the stability of P4 over the age of the solar system.

Standard image High-resolution image

The dependence of crossing time on the mass, m, of Nix and Hydra is well described by a power law

Equation (8)

The best-fit indices are γ ≈ −3.6, −4.3, −4.4,  and  − 4.6 for the bottom to top curves in Figure 3. The larger scatter about the 90th percentile "core" power law is partly due to Poisson noise, as fewer orbits contribute to this restrictive sample. Physical scatter caused by the shifting locations of the most stable regions could also contribute.

No known theoretical explanation exists for such a power-law scaling. Indeed, if crossing times scale with Hill radius as in Equation (6), the mass dependence is not a simple power law—instead the local power-law index would steepen toward lower masses. However, Duncan & Lissauer (1997) found that orbit crossing timescales for the Uranian moons also exhibit a power-law dependence on satellite mass.

Extrapolation along these power laws allows us to speculate about the stability of P4 for very low masses of Nix and Hydra. While such extrapolation is inherently uncertain, low masses are much more costly to simulate directly due to longer crossing timescales. The goal of extrapolation is to estimate the "allowed" masses of Nix and Hydra, where allowed means that P4 crossing timescales exceed the age of the solar system.

Including all sampled P4 orbits (the blue curve in Figure 3), the allowed masses of Nix and Hydra are implausibly low—below the minimum set by A = 1. However for low eccentricity P4 orbits (all other curves), a range of plausible Nix and Hydra masses are allowed. From extrapolation of the 90th percentile "core" sample (red curve), the lowest allowed albedo is A ∼ 0.5. Lower albedos are possible for special orbits, including the "most circular" orbits that we present next.

4.3. Stability of the "Most Circular" Orbits

The greater stability of low eccentricity P4 orbits shown above motivates the use of the "most circular" orbits about PC as an initial condition. These special orbits are not simply found by setting the Keplerian e = 0. Appendix A describes techniques to find these orbits.

We integrated trial orbits for 48 distinct P4 periods: 41 are uniformly spaced with period ratios (relative to Charon) from 4.79 to 5.31 and the remaining seven provide more dense sampling near the 5:1 resonance. At each orbital period 25 trial orbits were integrated. These 25 orbits are shifted in orbital phase along the most circular orbit by one Charon period each, as described in Appendix A.2.

Figure 4 shows the crossing times of the most circular orbits, for all orbital periods and for different Nix and Hydra masses. The median crossing times at each period are given by square symbols, with error bars spanning the 10th and 90th percentile. Lower limits are given where simulations did not run long enough to determine a timescale.

Figure 4.

Figure 4. Crossing times vs. orbital period for P4 when initialized on the "most circular" orbits about Pluto–Charon. The masses of Nix and Hydra (taking the same values as Figures 2 and 3) decrease from the bottom to top sets of colored symbols. Markers with error bars denote the median timescale and the 10th–90th percentile. Triangles indicate lower limits. Nominal locations of mean motion resonances with Charon, Nix, and Hydra are shown along the bottom axis. For lower masses, a narrow instability strip exists at the 5:1 commensurability with Charon. The gray horizontal error bar shows the observed mean motion of P4 (S11).

Standard image High-resolution image

Crossing times become longer as mass decreases, and the dependence on orbital period (or a) is quite complex. This general behavior agrees with the Keplerian initial conditions (Section 4.2). Figure 4 shows further that the period dependence is more varied and irregular for lower masses of Nix and Hydra. While it is difficult to understand all these fluctuations in detail, the trend is intuitively consistent with weaker resonant forcing being more narrowly focused at specific orbital periods.

Figure 4 shows a dramatic drop in crossing times near Charon's 5:1 resonance (C5:1). As also seen in Figure 2, this instability strip appears for lower masses of Nix and Hydra. At higher masses, Nix and Hydra perturbations wash out the influence of C5:1. Longer-lived P4 orbits exist both outside and inside the narrow instability strip at C5:1 Within the observational constraints, shown by the horizontal error bar, our analysis favors locations outside C5:1.

Figure 5 plots crossing times versus the mass of Nix and Hydra, with power-law fits overplotted. We restrict the range of periods to the observational constraint (S11), but with double the uncertainty for inclusiveness. The statistical measures of tc include: (1) "median," which takes the median of the values given by the symbols in Figure 4 (themselves median values over phases); (2) "max median," the longest phase-median tc, i.e., highest symbol; (3) "max 90%," the longest 90th percentile tc, i.e., highest upper error bar; and finally (4) "longest," simply the longest tc at any phase or period considered.

Figure 5.

Figure 5. Similar to Figure 3, but for crossing times, tc, of the "most circular" initial orbits with period ratios (to Charon) between 4.93 and 5.13. Blue squares give the median over both period and initial phase. Green squares (and red squares) give the longest of the median (and 90th percentile) tc at each period. Yellow squares give the longest tc of all orbits. For reference, the solid line gives power-law fit to the median tc for the "core" sample of Keplerian initial conditions. The most circular orbits give longer tc and thereby allow larger Nix and Hydra masses.

Standard image High-resolution image

For reference, the median tc for the relatively stable "core" sample of initial Keplerian parameters is overplotted. Compared to this reference case, the crossing timescales for the most circular orbits are significantly longer, especially toward the (more realistic) lower masses of Nix and Hydra. Longer crossing times equate to higher allowed masses (and lower albedos) for Nix and Hydra.

Extrapolating along the power-law fits in Figure 5 shows that A ≳ 0.3 is needed to achieve tc > 4 Gyr. This limit is more inclusive than the A ≳ 0.5 found in Section 4.2.2. We cannot definitely rule out even lower albedos. As already discussed, extrapolation could prove misleading.

Characterizing the most stable orbit is difficult. We do not base our estimate on the absolute longest lived orbits, which loosen our constraints. As shown in Figure 5, the longest tc's do not follow a simple power law and are therefore unreliable for extrapolation. Even without extrapolation, the longest tc's are highly subject to sampling, especially considering the pronounced period and phase dependence shown in Figure 4. It is also unclear whether P4 is likely to inhabit the most stable orbits, especially if those orbits occupy a tiny volume of phase space. Small neglected effects (such as collisions, see Section 3.4) could easily remove P4 from narrow pockets of parameter space. Thus, we base our constraints not on the absolutely most stable orbit, but on an average of orbits among the most stable.

4.4. Parameter Tests: Satellite Mass Ratios and Binary Eccentricity

We include a preliminary investigation of the effect of Nix and Hydra's mass ratio on the orbital stability of P4. Our main set of simulations fixed MNix/MHyd = 0.575. Here, we also investigate MNix/MHyd = 0.3 and 1.2, i.e., roughly half and double the fiducial value. For each mass ratio, we consider four values of the fixed total mass equal to the four highest mass cases in our main set of simulations. As in Section 4.3, we initialize the P4 orbits on the most circular orbits, with the same sampling of orbital periods and phases.

Figure 6 plots P4 lifetimes for these runs with different mass ratios. The dependence of orbital stability on the Nix/Hydra mass ratio is relatively minor compared to the stronger dependencies on Nix and Hydra's total mass and P4's orbital period. The decrease in orbital stability near the 5:1 commensurability—probably the most relevant feature—is evident in the lower mass runs for all mass ratios investigated, as are several other trends with period. The most prominent effect of the mass ratio is an increase in P4 crossing times at longer periods (≳ 5.1PCharon) for larger mass ratios, i.e., smaller Hydra masses. This behavior is unsurprising, and is qualitatively explained by three-body interactions. As shown in Figure 1, a more massive Hydra strongly perturbs orbits exterior to P4's measured orbit.

Figure 6.

Figure 6. Similar to Figure 4, but varying the mass ratio of Nix to Hydra (0.3, 0.575, and 1.2 for left, center, and right panels) for four values of the combined Nix and Hydra mass (with the same values and color scale as Figure 4.). The qualitative stability behavior is similar, including the decreased stability near 5:1 with Charon for the two lowest total masses (cyan and yellow data). At large periods, stability times increase for higher mass ratios, i.e., relatively smaller Hydra masses.

Standard image High-resolution image

For the two lowest total mass runs, the variation in median crossing timescales was ≲ 50% between the different mass ratios, compared to the fiducial case. Our preliminary investigation suggests that it is difficult for stability studies to precisely constrain Nix and Hydra's mass ratio at present. However, ever-improving orbit determinations should help, both on their own and in combination with stability studies.

We also investigated the effect of setting Charon's eccentricity to zero. We conducted this test with the methods of Section 4.2, i.e., uniform sampling of Keplerian osculating elements. As above, we investigated the four highest mass satellite cases. We find that reducing eC to zero has a modest stabilizing effect. The median crossing time increased by <50% (40%, 30%, 17%, and 42% for the highest to lowest satellite masses considered). Future work should aim to develop a more systematic understanding of how binary eccentricity, and other orbital parameters, affects stability.

5. SUMMARY AND DISCUSSION

We study the long-term stability of P4, the temporary name for the moon orbiting the PC binary between Nix and Hydra (S11). Our numerical integrations constrain both the orbit of P4 and the masses of Nix and Hydra. These constraints are coupled, so improved determination of P4's orbit will help refine the masses of Nix and Hydra and vice versa. We summarize our main results as follows:

  • 1.  
    Low eccentricity orbits of P4 are significantly more stable. Our integrations strongly disfavor P4 orbits with e > 0.02, as shown in Figures 2 and 3.
  • 2.  
    Period ratios (of P4 to Charon) between 4.98 and 5.01 are unstable on short timescales. Slightly larger or smaller period ratios are significantly more stable, as shown in Figure 4. Combined with the observed mean motion (S11), our results favor orbits just outside the 5:1 commensurability with Charon.
  • 3.  
    Even the most stable P4 orbits only survive if Nix and Hydra are sufficiently low in mass. We estimate MNix ≲ 5 × 1016 kg and MHyd  ≲  9 × 1016 kg are required for the stability of P4 over the age of the solar system. This constraint holds the mass ratio between Nix and Hydra fixed at the value implied by their mean brightness.
  • 4.  
    The albedos of Nix and Hydra are correspondingly constrained to A  ≳  0.3, assuming an internal density of 1 g cm−3. Higher density rocky bodies would require even higher albedos.
  • 5.  
    The above mass and albedo constraints rely on extrapolation of simulations with higher masses, as shown in Figure 5. Direct simulations alone disfavor A ≲ 0.16, for which the orbit crossing time of P4 is ≲ 107 yr.

Our mass limits based on the stability of P4 are a factor of 20 and 10 lower than the (1σ) astrometric upper limits of T08. The rendezvous of the New Horizons satellite with the Pluto system in 2015 July as well as ongoing Hubble astrometry (Buie et al. 2012) should greatly improve astrometric mass constraints. Neglecting P4, Beauvalet et al. (2012) combine current data with simulated New Horizons observations to show that mass errors on Hydra will be reduced to ∼4 × 1016 kg. This limit is already small enough to test our predictions. Hopefully, the inclusion of P4 will further tighten astrometric mass constraints. Ultimately, combining astrometry with long-term stability should provide the tightest and most robust dynamical constraints.

Our results generally support the leading model for the origin of the Pluto system: a giant impact that produces the PC binary (McKinnon 1989; Canup 2005) and the debris that forms its coplanar moons (Stern et al. 2006). The low eccentricity of P4 requires collisional damping. The inferred high albedo of Nix and Hydra is consistent with these being icy bodies. In the model of Canup (2011), the collision of differentiated bodies (or rock with icy mantles) forms Pluto and Charon plus a pure ice debris disk from which the moons can accumulate. Collisional stripping of an icy mantle similarly explains many properties of the dwarf planet Haumea: rapid rotation, collisional family members (Brown et al. 2007), and two high albedo icy moons (Ragozzine & Brown 2009). A possible weakness of the collisional scenario is that the debris disk forms much closer to PC than Nix's current orbit. Outward orbital migration has been proposed, but issues regarding simultaneous migration of multiple moons—which are likely more severe with P4—remain unresolved (Ward & Canup 2006; Lithwick & Wu 2008a; Peale et al. 2011).

The capture scenario for the system is implausible, especially for the circular and coplanar minor moons. However, the gravitational collapse scenario for the formation of lower mass Kuiper Belt binaries might also apply to the PC system (Nesvorný et al. 2010). This binary formation model is a natural extension of a leading planetesimal formation theory: solid debris accumulates via the streaming instability (Youdin & Goodman 2005; Youdin & Johansen 2007) and subsequently undergoes gravitational collapse (Youdin & Shu 2002; Johansen et al. 2009). Rotational fission is the extra twist needed for binaries. A particularly massive clump would be required for the formation of Pluto and Charon. However, massive clumps form by merging in streaming instability simulations (Johansen & Youdin 2007) and massive particle rings could form via related secular gravitational instabilities (Youdin 2011a; Shariff & Cuzzi 2011).

To explain Pluto's outer moons, some of the collapsing material must accumulate in a disk around the binary. Indeed, a possible benefit of the scenario is that higher angular momentum collapsing material could more readily produce distant moons. While plausible, this specific scenario has not been modeled in detail.

5.1. Pluto–Charon as Exoplanet Host Stars

As described in Section 3, the dynamics of the PC system are broadly relevant to our understanding of circumbinary dynamics. We conclude with a brief comparison to the circumbinary planets discovered by Kepler, and emphasize that Pluto is a guide to the circumbinary multi-planet systems that have yet to be discovered.

To make an analogy with planetary systems, we may scale the mass of Pluto to a solar mass. Charon then equates to a 0.12 M red dwarf. Nix, Hydra, and P4 would then have scaled minimum masses of ∼2MMars, ∼3.5MMars, and ∼3M, respectively. Kepler 16, 34, and 35 contain planets with masses of 1.1, 0.7, and 0.4 times Saturn (respectively) around binaries of secondary-to-primary mass ratios of 0.3, 0.97, and 0.91, respectively (Welsh et al. 2012). The period ratios of the planets to the inner binaries are 5.6, 10.4, and 6.3 (respectively). Unlike the Pluto system, there is no preference for being near n:1 resonances. However, the Kepler multiples around single stars do show a preference (as yet unexplained) for being near low-order mean motion resonances (Fabrycky et al. 2012).

Kepler's circumbinary planets are relatively close to the circumbinary stability boundary (HW). The period ratio of Kepler 16b is only 14% beyond this limit (Welsh et al. 2012) compared with 41% for Nix. Of course, the observational biases of transit surveys strongly favor shorter period planets (Youdin 2011b).

Kepler has revealed that compact multi-planet systems are common around single stars (Lissauer et al. 2011). However for circumbinary systems, a planet too close to the stability boundary is unlikely to have other planets nearby. Thus, circumbinary planets detected at a safer distance from the stability boundary are more likely to be in circumbinary multi-planet systems—the true Pluto analogs.

Our study shows that multi-planet scattering is more violent around a binary than a single star. Thus, orbital stability will require larger spacing between circumbinary planets, making detection of multi-planet systems more challenging. Nevertheless, Pluto's rich set of companions bodes well for a fruitful search.

We thank Matt Holman, Alex Parker, Dan Fabrycky, and Josh Carter for helpful discussions. A.N.Y. and K.M.K. thank Tom Aldcroft, Tom Robitaille, Brian Refsdal, and Gus Muench for http://python4astronomers.github.com. Portions of this project were supported by the NASA Astrophysics Theory Program and Origins of Solar Systems Program through grant NNX10AF35G and the Outer Planets Program through grant NNX11AM37G. The computations in this paper were run on the Odyssey cluster supported by the FAS Science Division Research Computing Group at Harvard University.

APPENDIX A: "MOST CIRCULAR" ORBITS ABOUT A BINARY

Section 4.3 presents integrations of the Pluto system with P4 initially placed on the most circular orbit about the PC binary. Lee & Peale (2006) develop an analytic theory for the orbits of test masses about Pluto and Charon (or more generally any circular binary). Their theory separates orbital motion into three components: (1) circular motion of the guiding center about the barycenter, (2) oscillations forced by binary motion, and (3) the free or epicyclic eccentricity. Setting the epicyclic eccentricity (distinct from Keplerian eccentricity) to zero gives the most circular orbits. Their theory applies to a circular binary orbit, for which the potential is constant in a rotating frame. To account for the small (but probably overestimated) eccentricity of Charon in the T08 orbit solution, we instead use the invariant loop method of Pichardo et al. (2005), which Lithwick & Wu (2008a) applied to the Pluto system. We demonstrate below that the adopted value of Charon's eccentricity gives only a minor correction to the Lee & Peale (2006) solutions for the most circular orbits.

A.1. Finding Invariant Loops

In time-periodic potentials—such as an eccentric binary—special orbits form a closed loop when plotted stroboscopically, i.e., when a snapshot is taken once per orbit of the potential. For concreteness, we take this snapshot when the binary orbit is at periapse. These special, stroboscopically closed orbits are the most circular orbits. In the case of resonant orbits, the stroboscopically closed loop breaks into a chain of islands.

Invariant loops in the plane of the binary can be found by iterative methods. Integrations start when the binary is at periapse and the test particle is aligned with the binary. By symmetry, the radial velocity is zero at this location for an invariant loop. Thus at a given radial distance, the only initial condition to vary is the azimuthal velocity, vo. For the correct choice of vo, the stroboscopic orbit forms a one-dimensional curve. To iterate toward this orbit, we vary vo to minimize the radial dispersion of the stroboscopic orbit.

Lithwick & Wu (2008a) measured the radial dispersion in narrow azimuthal bins so that the invariant loop has a nearly constant radial distance in the bin. We found faster and more reliable convergence by fitting the stroboscopic orbits to a (fourth order) cosine series, and measuring the radial dispersion about that best-fit orbit.

A.2. Implementation as Initial Conditions

In Section 4.3, we present integrations where Nix and Hydra perturb P4 from these most circular orbits. The initial orbital phase along the most circular orbit must be consistent with the initial orbital phase of the binary. This initial phase is not unique as P4 can be advanced along the orbit by an integer number of Charon periods. For different orbital phases, the position of P4 changes relative to Nix and Hydra. Due to the sensitive dependence on initial conditions, stability timescales also change. Thus, our simulations sample many (typically 25) initial phases of P4 for every trial orbital period.

A.3. Properties of the Most Circular Orbits

Figure 7 plots the time evolution of the most circular test particle orbits in the vicinity of P4. These orbits only include the perturbations from the PC binary, and not those due to Nix or Hydra. Each curve represents a different orbital period, ranging from 4.8 to 5.2 PC periods. None of these orbits are in a 5:1 resonance with Charon, and the effect of the 5:1 is negligible due to the low eC, unlike the study of the 4:1 for a high eC by Lithwick & Wu (2008a).

Figure 7.

Figure 7. Evolution of the "most circular" test particle orbits about Pluto–Charon vs. time. The orbits are in the vicinity of P4 and neglect perturbations from Nix and Hydra. Left: radial distance from the Pluto–Charon barycenter. The denser packing of lines near the middle of the plot is from finer sampling near the 5:1 resonance. Center and right: Keplerian semimajor axis and eccentricity (measured from the Pluto–Charon barycenter). The oscillation of these elements demonstrates the non-Keplerian nature of orbits about the Pluto–Charon binary.

Standard image High-resolution image

The left panel plots radial distance from the PC barycenter. The non-circularity, given by the magnitude of the radial excursions relative to the mean r, is Δr/(2r) ∼ 2 × 10−3. The periodic radial oscillations have well-understood timescales (Lee & Peale 2006). The faster oscillations correspond to the synodic period with Charon (∼5/4 Charon periods) and the first harmonic at half that period. The slower oscillations are on the epicyclic period (∼5 Charon periods) of the test particle itself. The most circular orbits about a circular binary do not have these longer period epicyclic oscillations (Lee & Peale 2006). The epicyclic oscillations are only modestly larger than the synodic timescale variations. When Nix and Hydra are introduced to numerical integrations they rapidly excite a larger epicyclic eccentricity. Thus, Charon's adopted eccentricity is not expected to strongly affect orbital stability. The integrations in Section 4.4 affirm this expectation.

The middle and right panels of Figure 7 plot the osculating (instantaneous) Keplerian a and e, respectively, about the barycenter to demonstrate the non-Keplerian nature of these orbits. The Keplerian a is both systematically larger than the cylindrical radius and varies more significantly with time. The Keplerian e oscillates and reaches values ∼0.015, much larger than the actual radial excursions. This plot demonstrates clearly that setting the osculating e = 0 does not give the most circular orbit, and why e had to be increased above ∼0.01 in Figure 2 to noticeably affect orbital stability.

APPENDIX B: CODE DETAILS

We use the Radau integrator (RA15) included with Swifter because it provides a good combination of simplicity, accuracy and efficiency. Standard symplectic integrators are not appropriate for this problem because they assume a dominant central mass. Modified symplectic integrators that handle an inner binary have been developed (Chambers et al. 2002; Beust 2003). However for this study, we use non-symplectic integrators that make no assumptions about the masses and orbital architecture of the system. We also tested the standard Bulirsch–Stoer (BS) integrator, another non-symplectic method. We opted for the higher order RA15 method because we do not need to handle close encounters. Once orbit crossing occurs the system is deemed unstable.

The main problem we encountered using non-symplectic methods (RA15 and BS) is that they are both inherently variable time step algorithms that the Swifter wrapper occasionally forces to use a fixed time step in order to produce output at regular intervals. (Mercury behaves similarly.) Energy conservation suffers if these forced time steps are too frequent. We were able to mitigate this problem with suitable choices of the output interval, the user-provided dt, which we measure in PC orbital periods.

If dt is very short compared to the optimal time step determined by the RA15 algorithm, energy conservation is good, but the algorithm's efficiency suffers. At the other extreme, if dt is much longer than optimum RA15 time step, then forced non-optimal time steps are rare, and energy conservation suffers little. The main cost is that one does not have finely sampled output for analysis. For intermediate dt, many non-optimal time steps can degrade energy conservation significantly. Quantitatively, both long (dt = 1000 PC periods) and short (dt < 0.01 PC periods) output times conserved energy to ΔE/E ≈ 10−12 over 105 PC periods. For comparison, energy conservation drops to ΔE/E ≈ 10−9 for dt = 0.1 PC periods.

For all long integrations, we use a time step of dt = 1000 PC periods. We set the user-supplied error tolerance parameter to 10−12 for all integrations. The energy error accumulation appears to be, at worst, linear in time. Extrapolating to our longest runs, 109 PC orbits, we expect a total energy error of roughly 1 part in 108.

We hope that our description of this issue is of some use for those using non-symplectic codes, especially for long timescale integrations where energy conservation is desired.

Footnotes

  • While the 1σ errors on the mass of Nix do not quite extend to zero mass, T08 discuss the difficulty of error estimation in their high dimensionality fits.

  • While the best-fit T08 masses have a different ratio, large uncertainties mean the ratio has little significance. Furthermore, integrations showed that the T08 masses destabilize P4 orbits on a median timescale of ≲ 103 yr.

  • We use Jacobian elements measured relative to the barycenter of Pluto, Charon, and Nix.

Please wait… references are loading.
10.1088/0004-637X/755/1/17